Next Article in Journal
Current Applications of Nanoemulsions in Cancer Therapeutics
Previous Article in Journal
Sesquiterpenoids from Tussilago farfara Flower Bud Extract for the Eco-Friendly Synthesis of Silver and Gold Nanoparticles Possessing Antibacterial and Anticancer Activities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic Attributes of NiFe2O4 Nanoparticles: Influence of Dysprosium Ions (Dy3+) Substitution

1
Department of Biophysics, Institute for Research & Medical Consultations (IRMC), Imam Abdulrahman Bin Faisal University, P.O. Box 1982, Dammam 31441, Saudi Arabia
2
Department of Physics, College of Science, Imam Abdulrahman Bin Faisal University, P.O. Box 1982, Dammam 31441, Saudi Arabia
3
Department of Physics, Hitit University, Çevre Yolu Bulvarı-Çorum 19030, Turkey
4
Mechanical and Energy Engineering Department College of Engineering, Imam Abdulrahman bin Faisal University, P.O. Box 1982, Dammam 31441, Saudi Arabia
5
Department of Chemistry, Bharath Institute of Higher Education and Research, Bharath University, Chennai, Tamil Nadu 600073, India
6
Department of Nano-Medicine Research, Institute for Research & Medical Consultations (IRMC), Imam Abdulrahman Bin Faisal University, P.O. Box 1982, Dammam 31441, Saudi Arabia
7
Scientific-Practical Materials Research Centre NAS of Belarus, 19 P. Brovki Street, 220072 Minsk, Belarus
8
Department of Electronic Materials Technology, National University of Science and Technology MISiS, Leninsky Prospekt, 4, Moscow 119049, Russia
9
Laboratory of Crystal Growth, South Ural State University, Lenin Prospect, 76, Chelyabinsk 454080, Russia
*
Authors to whom correspondence should be addressed.
Nanomaterials 2019, 9(6), 820; https://doi.org/10.3390/nano9060820
Submission received: 2 May 2019 / Revised: 22 May 2019 / Accepted: 28 May 2019 / Published: 31 May 2019

Abstract

:
This paper reports the influence of dysprosium ion (Dy3+) substitution on the structural and magnetic properties of NiDyxFe2−xO4 (0.0 ≤ x ≤ 0.1) nanoparticles (NPs) prepared using a hydrothermal method. The structure and morphology of the as-synthesized NPs were characterized via X-ray diffraction (XRD), scanning and transmission electron microscope (SEM, and TEM) analyses. 57Fe Mössbauer spectra were recorded to determine the Dy3+ content dependent variation in the line width, isomer shift, quadrupole splitting, and hyperfine magnetic fields. Furthermore, the magnetic properties of the prepared NPs were also investigated by zero-field cooled (ZFC) and field cooled (FC) magnetizations and AC susceptibility measurements. The MZFC (T) results showed a blocking temperature (TB). Below TB, the products behave as ferromagnetic (FM) and act superparamagnetic (SPM) above TB. The MFC (T) curves indicated the existence of super-spin glass (SSG) behavior below Ts (spin-glass freezing temperature). The AC susceptibility measurements confirmed the existence of the two transition temperatures (i.e., TB and Ts). Numerous models, e.g., Neel–Arrhenius (N–A), Vogel–Fulcher (V–F), and critical slowing down (CSD), were used to investigate the dynamics of the systems. It was found that the Dy substitution enhanced the magnetic interactions.

1. Introduction

Lately, nanosized Ni-spinel ferrites have widely been used in magnetic storage media, magnetic adsorbents, telecommunication, catalysts, microwave absorbers, and computer memories [1,2,3,4,5]. Further, the structural and magnetic characteristics of nanosized Ni-spinel ferrites have been enhanced by substituting varieties of magnetic, nonmagnetic, and rare earths ions [6,7,8]. Such improvement was majorly attributed to the distribution of substitution ions between tetrahedral and octahedral sites of the host and eventual influence on their magnetic moments [8]. Moreover, the use of rare earths as substitution agents in the spinel ferrite was found to cause structural distortion through symmetry preserving the lattice strain and thereby modifying the overall properties of nanosized spinel ferrites [9]. Intense research efforts have been made to enhance the structure, optical, and magnetic properties of rare earth ion substituted nanosized Ni-spinel ferrites [2,3]. Meanwhile, different theories and models have been applied to achieve a better understanding of the unique magnetic properties of these doped nanosized Ni-spinel ferrites [2,3]. Modified magnetic properties of Erbium (Er3+) and Samarium (Sm3+) substituted nickel ferrites were analyzed by low temperature magnetization, and zero-field cooled and field cooled (ZFC-FC) measurements. Such samples revealed a superparamagnetic (SPM) behavior with very low coercivity (Hc) and remanence magnetization (Mr), making them suitable for developing soft magnets [10]. A substitution of Pr, Sm, and La in Ni0.5Zn0.5Fe1.95R0.05O4 nanoparticles (NPs) was found to modify their magnetic and dielectric properties [11]. Conversely, Cerium (Ce) substituted Ni-Zn at low contents (0.0 ≤ x ≤ 0.1) produced high saturation magnetization (Ms), remanence, and coercivity [12].
So far, studies on Dy3+ substituted nanosized Ni spinel ferrites are rare. J. Sahariya et al. [13] reported the temperature dependent spin momentum densities of NiFe2−xRExO4 (x = 0, 0.05 and RE = Dy, Gd) ferrites measured by using a magnetic Compton spectrometer. The experimental profiles of NiFe2O4 with doping of Dy and Gd show a similar spin moment to the non-doped sample. The contribution of different constituents in the formation of total spin moment is deduced from the analysis of the Compton line shape. It is seen that Dy3+ or Gd3+ doping ions at Fe3+ sites lead to a redistribution of the spin moment at Fe3+ and RE3+ sites. The temperature dependent magnetic Compton profiles are decomposed into the constituent profiles of Ni, Fe, Dy/Gd, and diffuse components, arising due to O-2sp states. A decrease in the Fe spin moment from 0.55 ± 0.03 μB/f.u. (in NiFe2O4) to 0.41(0.50) ± 0.03 μB/f.u. is observed on the partial substitution of Dy(Gd). K. Kamala Bharathi et al. [14] reported a correlation between the microstructure, electrical, and optical properties of Dysprosium-doped nickel-ferrite (NiFe1.925Dy0.075O4) thin films fabricated using sputter-deposition using a stoichiometric bulk target prepared by tsolid-state chemical reaction. Recently, M.A. Almessiere et al. [15] investigated the effects of dysprosium substitution on the structural, microstructural, and magnetic properties of NiFe2O4 nanoparticles. The formation of the cubic phase of Ni nanosized ferrites was confirmed. The magnetic properties were done by analyzing measurements of the magnetization versus applied field M(H). These measurements were carried out at two different temperatures—room temperature (T = 300 K) and low temperature (T = 10 K). A noticeable improvement in the differently deduced magnetic parameters, including saturation magnetization (Ms), remanence (Mr), and coercivity (Hc), was observed at both room (T = 300 K) and low (T = 10 K) temperatures with Dy substitution. The increase in the different magnetic parameters is mainly attributed to the strengthening of A–B exchange interactions owing to the substitution of Fe3+ ions with Dy3+ ions with larger ionic radii, the formation of local strains, and the increase in the magnetic moments (NB).
To the best of our knowledge, there is no study reporting the Mossbauer, zero-field cooled (ZFC), and field cooled (FC) magnetizations and AC susceptibility measurements for Dy substituted nanoparticles of Ni spinel ferrites. Accordingly, we examined in the present study the influence of Dy3+ doping on the structure, morphology, Mossbauer, ZFC-FC magnetization, and AC susceptibility of NiDyxFe2-xO4 (0.0 ≤ x ≤ 0.1) nanoparticles. Various products were synthesized using a hydrothermal process.

2. Experimental

Various specimens with the chemical formula NiDyxFe2−xO4 (0.0 ≤ x ≤ 0.1) have been prepared by a hydrothermal process. Analytical grade high purity chemical reagents including Dysprosium III nitrate hydrate (Dy (NO3)3 H2O), Nickel nitrate (Ni(NO3)2 6H2O), Iron III nitrate (Fe(NO3)3 9H2O), and Sodium Hydroxide (NaOH) were taken as starting materials. Then, the stoichiometric amount of these compositions was mixed in distilled water and the resulting mixture was left at room temperature. Next, a clear solution was made for pH adjustment using 2M of NaOH solution and placed in a stainless steel-Teflon autoclave to initiate the hydrothermal treatment. Finally, the obtained sample was dried and grinded for further experimental analyses.
A Rigaku Benchtop Miniflex X-ray diffraction (XRD) diffractometer (Tokyo, Japan) with Cu Kα radiation at room temperature (RT) over a 2θ range from 20° to 70° was used for the structural analysis. Scanning electron microscopy (SEM, FEI Titan 80–300 kV FEG S/TEM, Hillsboro, OR, USA), along with energy dispersive X-ray spectroscopy (EDX) and transmission electron microscopy (TEM; FEI, Morgagni 268, Prague, Czech Republic), was used for the morphological and composition analyses. AC magnetic susceptibility, zero-field cooled (ZFC), and field cooled (FC) magnetization measurements were performed using a superconducting quantum interference device (PPMS DynaCool, Quantum Design, San Diego, CA, USA). ZFC-FC magnetizations were carried out at a temperature range of 2–400 K under a DC field of 100 Oe. Real and imaginary parts of AC susceptibility measurements were done at temperatures ranging from 350 to 2 K in the presence of an ac applied magnetic field of H ac = 10 Oe and in different frequencies ranging from 50 to 10,000 Hz. Mössbauer spectra were performed at room temperature using a conventional Mössbauer spectrometer (Fast Com Tec PC-moss II, Oberhaching, Germany) under the constant acceleration mode using 57Fe in an Rh matrix with an approximate activity of 10 m Ci. The speed scale and the velocity were calibrated using α-Fe and laser interferometry, respectively. The recorded spectra were analyzed and fit to the inbuilt Win-Normos fitting software (WISSEL company, Duisburg, Germany).

3. Results and Discussion

3.1. Structure

Figure 1 displays the XRD powder patterns of the studied NiDyxFe2−xO4 (0.0 ≤ x ≤ 0.1) nanoparticles. Irrespective of Dy3+ contents, all the XRD peaks have been verified to the cubic nanosized Ni-spinel ferrites (JCPDS Card Number 54-0964), indicating the lattice site compatibility of the Dy3+ ions in the Ni spinel ferrite structure in the absence of any impurity. The calculated structural parameters of the prepared nanosized spinel ferrites were enlists in Table 1. It was realized that lattice parameters elongate due to the increase in the substitution contents and also show ionic radii disparity between Fe3+ (0.78 Å) and Dy3+ (1.03 Å). The calculated crystallite sizes (DXRD) were in the range of 24 to 35 nm [16].

3.2. Morphology

Figure 2 displays the SEM pictures, EDX spectra, and elemental mapping results for x = 0.01 and 0.05 samples. The surface morphology of these samples revealed a high degree of nanoparticle agglomeration. EDX analysis proved the existence of appropriate elements (Fe, Ni, Dy, and O) in the samples. Figure 3 illustrates the TEM image of the nanosized spinel ferrites containing Dy3+ contents of 0.05. The inset (Figure 3) displays the selected angle electron diffraction (SAED) patterns of the corresponding sample that confirmed the agglomeration of single crystalline Ni spinel ferrite nanoparticles.

3.3. Mössbauer Spectra

Figure 4 depicts the Mössbauer spectra of synthesized NiDyxFe2−xO4 (0.0 ≤ x ≤ 0.1) NPs at room temperature. Table 2 enlists various Mössbauer parameters calculated by spectral fitting using three sextets (A for the tetrahedral sites and B and B1 for the octahedral sites). Fe3+ ions in the tetrahedral A site are characterized by a large hyperfine field with an insignificant isomer shift. Conversely, the other two sextets with a comparatively smaller hyperfine field signify the occupation of Fe3+ at two dissimilar environments in the B-site [17]. Besides the ferromagnetic sextets, a minute paramagnetic doublet with quadrupole-splitting was evidenced for NiDy0.01Fe1.99O4, NiDy0.07Fe1.93O4, and NiDy0.1Fe1.9O4 NPs. The occurrence of such a paramagnetic doublet was attributed to the fractions of Fe3+ with fewer nearest neighbors that possessed magnetically ordered spins. Interestingly, in the spinel ferrite structure, Fe3+ did not contribute to the super exchange interaction [18].
The achieved relative area for the A and B sites clearly indicated the occupation of Ni3+ in the A and B sites. Kumar et al. acknowledged the preferential occupation of Dy3+ in the octahedral B sites of Co-ferrites [19,20]. Thus, the cation distribution in the proposed Dy3+ substituted nanoferrites was obtained following the formula unit of (NiyFe1−y)A (Ni1−yDyxFe1+yx)B. The distribution of Fe3+ over the A and B sites was observed to be relative to the proportional area of A and B in the Mossbauer sub-spectra. Table 2 summarizes the approximate cation distribution obtained from the Mössbauer spectra. The results in Table 2 revealed that Fe3+ cations emigrated from the B site to A site due to Dy3+ substitution. The line width of the A site was randomly altered, whereas for B and B1 sites, it was enhanced with the substitution of Fe3+ (0.64 Å) in the B sites by Dy3+ (0.91Å) having larger ionic radii than the former one. This observation authenticated the increase in the degree of disorder due to the substitution on B sites.
The values of the hyperfine magnetic fields for the A and B sites (Table 2) in the studied nanosized spinel ferrites were first reduced with an increase in Dy3+ contents up to 0.05 and then enhanced at 0.07. Eventually, the hyperfine field for A site was continuously enhanced, but for the B site it was diminished. This alteration in the hyperfine field for the A and B sites was attributed to the addition of the diamagnetic Dy3+ that replaced the ferromagnetic Fe3+ with a higher magnetic moment (5 µB) and lowered the average number of magnetic linkages ( Fe A 3 + O Fe B 3 + ). Thus, Fe3+ nuclei experienced a reduction in the magnetic field at both the sublattices up to the Dy3+ content of 0.05. Beyond 0.05, the number of Fe3+ at the A site was augmented, thereby increasing the hyperfine magnetic field and magnetic moment of Fe3+ at the A site.

3.4. ZFC-FC Magnetizations

Figure 5 shows the curves of zero-field-cooled (ZFC) and field-cooled (FC) temperature dependencies of the magnetization, MZFC (T) and MFC (T), of NiFe2−xDyxO4 (where x = 0.00, 0.03 and 0.09) NPs. These measurements were performed in a temperature interval ranging between 2 and 400 K under a DC field of 100 Oe. For MZFC (T) measurements, the sample was cooled, first of all, from room temperature (RT) to a very low temperature in the absence of an applied field and subsequently the magnetization was recorded by increasing the temperature in the presence of the field. However, in the MFC (T) measurements, the magnetization was recorded by cooling the product in the presence of applied field. A splitting and a large irreversibility between MZF (T) and MFC (T) curves for different synthesized products can be clearly seen in Figure 5. The MFC (T) increased gradually and remained constant below temperature T s , while the MZFC (T) decreased with a lowering of the temperature down to about 4 K. The dispersion of MZFC-MFC versus T curves is congruent with the poly-disperse character of magnetic NPs, with a correlated distribution in particle size and individual anisotropy axes [21]. The enlargement could also owe to dipolar interactions among particles [21].
It is reported in the literature that the manifestation of a peak in the MZFC (T) plots is associated to the blocking temperature (TB) [22]. The curves of MZFC (T) of the prepared products showed an incomplete maximum or broad maximum at around the temperature noted TB. This is typical for superparamagnet (SPM) materials, which show the properties of classical paramagnet materials (PM) above TB, where the total spin is equal to the spin of a whole NPs but behave as ferromagnetic (FM) materials below their blocking temperature (TB). Below TB, the MFC (T) and MZFC (T) curves considerably diverge, and the various ferrite NPs are in the FM state (blocked state). Above TB, the MFC (T) and MZFC (T) curves coincide, which is because all NPs are in the same SPM state. At T = TB, the thermal activation overcomes the magnetic anisotropy barrier, which leads to fluctuations in magnetization [23]. Therefore, the wide peak at TB in the MZFC (T) curves is an indication of a broadened energy barrier distribution. Further, it can be seen that the blocking temperature varies by increasing Dy substitution content. The non-substituted product NiFe2O4 shows a blocking temperature around TB ≈ 390 K. The x = 0.03 product exhibits a well-defined TB at around 300 K. By further increasing the amount of Dy, the product synthesized with x = 0.09 was not able to reach the TB within 400 K, so the TB value is superior than 400 K for x = 0.09. It can be seen clearly that the blocking temperature decreases for a lower Dy content (x = 0.03) and then increases for higher content (x = 0.09). The dependence of the TB on particles size has been reported in previous studies [24]. The lower TB is attributed to smaller particle size or narrow size distribution. However, the higher TB represents a larger particle size. Nevertheless, the different x = 0.00, 0.03 and 0.09 products show approximately same particles size. Therefore, the variations in blocking temperature with substitution effects are not predominantly influenced by the grain size. Thus, in addition to the particle size effect, the TB could also be affected by numerous other extrinsic factors, mostly related to interactions among particles and intrinsic factors that principally include a magneto–crystalline, surface and shape anisotropy [22,25]. We noticed in the synthesized product with x = 0.03 that the MZFC (T) exhibits a breaking at temperatures indicated by the dashed circle in Figure 5.
On the other hand, the MFC (T) curves increase smoothly for different samples with a decrease of temperature, while a kind of saturation in the magnetization is noticed below the temperature noted by T s for all samples. It is reported in the literature that for SPM nanoparticles, the curve of MFC (T) increases continuously [26,27]. Nevertheless, in the case of super-spin glass (SSG) systems in which the interactions among particles are strong, a flat type or a slow increase is observed [26,27]. Therefore, the detected flat nature below T s in the MFC (T) curves establishes the occurrence of an SSG-like state. The origin of the observed magnetic features in ZFC-FC magnetization part will be discussed in detail in the analyses of the AC susceptibility measurements. The latter are a useful way to identify the freezing dynamics of the spin-glass (SG) materials.

3.5. AC Susceptibility

For measurements of AC susceptibility ( χ ac ), an AC magnetic field ( H ac ) is applied to the sample and, as a consequence, a resultant magnetic moment is measured. The χ ac is represented as follows:
χ ac = χ + i χ
where χ’ real and χ imaginary parts are, respectively, the in-phase and out-phase components of χ ac . It should be noted that the relaxation time ( τ ) of the AC susceptibility measurement is not based upon the energy barrier ( E a = K eff V where K eff is the effective anisotropy constant and V is the volume of particles). However, it is influenced by the external excitation frequency. The AC susceptibility measurements give important details about the dynamics of the systems and the strength of exchange interactions between the magnetic nano-particles (MNPs) and between the different cations.
Firstly, we will discuss the χ’ real part measurements of the two NPs samples with x = 0.00 and 0.03. Figure 6 presents the curves of χ’versus T, ranging from 350 to 2 K, for x = 0.00 and 0.03 products, performed in the presence of an H ac = 10 Oe and in a frequency range of 50–104 Hz. The magnitude of χ’ for x = 0.03 increased slightly compared to the non-substituted product (x = 0.00), which is in accordance with MZFC (T) and MFC (T) measurements. The in-phase AC susceptibility data of the different samples showed dispersion and a decrease in magnitude while increasing the applied frequency from 50 Hz to 10 kHz. The χ’(T) curve of the non-substituted NiFe2O4 NPs exhibited a peak at around 300 K. However, the x = 0.03 product did not show any peak up to 350 K.
Figure 7 shows the χ ( T ) curves for NiFe2-xDyxO4 (where x = 0.00 and 0.03) performed in a H ac = 10 Oe and in the frequency range of 50–104 Hz. It can be seen that both samples display two peaks—the first at the higher temperature indicated by TB in the figure, which is associated with magnetic blocking of huge core spins, and the second indicated by T s , which can be associated with the spin-glass freezing on the surface of a single NP [28]. The χ’(T) curves do not offer any information about these two peaks, hence from now on we will focus only on analyses of χ ( T ) curves.
At the same applied frequency, the TB and T s shifted to lower temperatures with Dy substitution compared to the non-substituted one. This is consistent with the MZFC analyses. Both the blocking temperature TB and spin-glass freezing temperature T s are affected by frequency. Both show a shift to higher temperatures upon increasing the value of the applied frequency ( f ). Similar behavior is observed in the spin-frustrated system of CoFe2O4 NPs dispersed in an SiO2 matrix [28]. The shifting with f is helpful for evaluating dynamic magnetic behaviors, deducing the anisotropic energy, the magnetic anisotropy, and the interaction strength between MNPs.
Various physical laws can be used to investigate the f-dependence shift of TB and T s temperatures. The Neel–Arrhenius (N–A) law was first tested to fit the experimental data (Figure 8). This theory is valid for thermal excitations of non-interacting single-barrier NPs and is expressed as follows [29,30]:
τ = τ o   e x p ( E a / k B T )
where τ = 1 / f is the measured time, τ 0 is the jump attempt time (in the range of 10−9–1013 s), k B is the Boltzmann constant, and E a = K eff V is the activation energy barrier. The estimated values of τ 0 = 1 / f 0 , E a / k B and K eff for different samples are given in Table 3. The best N–A fit offers very unreasonable values for τ 0 and E a / k B . This indicates that the synthesized products do not obey the thermally activated N–A law and, as a consequence, they are non-interacting.
The Vogel–Fulcher (V–F) law is a useful model for investigating the interactions between NPs. This law uses an additional parameter, T 0 , that represents the strength of inter-particle interactions. Based on this model, the relaxation is described as follows [29,30]:
τ = τ o   exp [ E a / k B ( T T 0 ) ] .
The fitting data using the V–F law of the plots of f vs. TB and f vs. T s for the prepared products are illustrated in Figure 9a,b, respectively. The different estimated parameters are summarized in Table 3. The analysis of f-dependent TB now gives reasonable τ 0 and E a / k B values. Obviously, the T0 values are not negligible compared to TB. The occurrence of T0 confirms the presence of moderate inter-particle interactions between the NPs [28,29,30]. It is found, moreover, that τ 0 increased more for the x = 0.03 product than for the x = 0.00 one. The increase in the τ 0 for x = 0.03 product suggests the strengthening of interactions between NPs [29,30]. Compared to the x = 0.00 product, the values of E a / k B and K eff improved with Dy substitution for x = 0.03. This improvement iresulted from the strengthening of magnetic interactions among different NPs and the increase of magnetic anisotropy sources [29,30].
In other hand, the investigation of f vs. T s provides unphysical values for τ 0 . Therefore, the critical slowing down (CSD) law is used to study the presence of SG behavior in the synthesized NPs. Based on this model, the relaxation is expressed as [30]:
τ = τ 0 * [ T s T g 1 ] ( z υ )
where τ 0 * is associated to the coherence time of coupled individual “atomic” spins in the NP (in the range 10−6–10−13 s) [31], T g is the SG freezing temperature, and T s is the f-dependent freezing temperature. The z υ is the critical exponent that offers information about the SG, and it varies from 4 to 12 for various SG systems [28]. We fit the same f vs. T s data using the CSD law, in order to examine the possibility of the SG nature (Figure 10). The various deduced parameters are listed in Table 3. The obtained reasonable values of τ 0 , T g and z υ proved the existence of SG behavior in the prepared samples. Similar comportment has been reported in numerous products, such as CoFe2O4/(SiO2)x systems [28], Fe3O4 MNPs ( z υ = 8.2 and τ   ~   10 9   s ) [32,33], soft ferrite Ni0.3Zn0.7Fe2O4 NPs ( z υ = 8.01 and τ   ~   10 12   s ) [34], and La0.9Sr0.1MnO3 NPs [35]. It is reported that the strength of magnetic interactions increases based on the decreasing z υ exponent. The non-substituted NiFe2O4 product exhibits a z υ value equal to 5.11, and it decreases to 3.95 with Dy substitution for x = 0.03. This result indicates the improvement of the magnetic interactions among NPs for the x = 0.03 product.

4. Conclusions

Series of Dy3+ doped ferrite nanoparticles of NiFe2-xDyxO4 (0.0 ≤ x ≤ 0.1) NPs were prepared using a hydrothermal method. The as-prepared specimens were thoroughly characterized using various analytical measurement techniques to determine their structure, morphology, and distinct magnetic traits. An XRD pattern of nanoferrites confirmed the formation of a spinel cubic structure. The Mössbauer spectra displayed the cation distribution, verifying the occupation of Dy3+ ions at octahedral B sites. The measurements of ZFC-FC magnetization and AC susceptibility were examined. The magnetic data showed the existence of two critical temperatures. The first one is the blocking temperature T B , which corresponds to the magnetic blocking of huge core spins, and the second is T s , which can be associated to spin-glass freezing arising on the surface of a single NP. The T B and T s temperatures vary with the Dy substitution. This is due to numerous intrinsic and extrinsic factors that principally include magneto-crystalline factors, interactions among particles, surface, and shape anisotropy. Furthermore, both the T B and T s temperatures are affected by applied frequency. Various physical laws, such as Neel-Arrhenius, Vogel-Fulcher, and critical slowing down, are used to investigate the f-dependence shift of the T B and T s temperatures. It was found that the Dy substitution enhances the magnetic interactions. Compared to the x = 0.00 product, the E a / k B and K eff values improved for x = 0.03. This proved that the magnetic interactions are enhanced due to the Dy substitution.

Author Contributions

Synthesis of the sample (M.A.A., Y.S. and A.B.). XRD investigation and analysis (M.A.A. and A.B.). Investigation and writing of the ZFC-FC magnetization and AC susceptibility (Y.S.). Investigation and analysis of the Mossbauer spectra (H.G.). Participation in drafting the article or revising it critically for important intellectual content (S.A., A.M., I.E. and A.V.T). Supervision, review, and editing (M.A.A., Y.S. and A.B.).

Funding

The authors highly acknowledge the Institute for Research & Medical Consultations (Projects application No. 2017-IRMC-S-3; No. 2018-IRMC-S-1 and No. 2018-IRMC-S-2) and the Deanship for Scientific Research (Projects application No. 2017-576-IRMC and No. 2018-209-IRMC) of Imam Abdulrahman Bin Faisal University (Saudi Arabia) for supporting this study.

Acknowledgments

The authors thank the Institute for Research & Medical Consultations (IRMC) of Imam Abdulrahman Bin Faisal University (IAU—Saudi Arabia) for the supports. The technical assistance provided by Core Labs of King Abdullah University of Science and Technology (KAUST) is highly appreciated.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kavas, H.; Kasapoğlu, N.; Baykal, A.; Köseoğlu, Y. Characterization of NiFe2O4 nanoparticles synthesized by various methods. Chem. Pap. 2009, 63, 450–455. [Google Scholar] [CrossRef]
  2. Ati, A.A.; Othaman, Z.; Samavati, A. Influence of cobalt on structural and magnetic properties of nickel ferrite nanoparticles. J. Mol. Struct. 2013, 1052, 177–182. [Google Scholar] [CrossRef]
  3. Vigneswari, T.; Raji, P. Structural and magnetic properties of calcium doped nickel ferrite nanoparticles by co-precipitation method. J. Mol. Struct. 2017, 1127, 515–521. [Google Scholar] [CrossRef]
  4. Wahba, A.M.; Mohamed, M.B. Structural, magnetic, and dielectric properties of nanocrystalline Cr-substituted Co0.8Ni0.2Fe2O4 ferrite. Ceram. Int. 2014, 40, 6127–6135. [Google Scholar] [CrossRef]
  5. Sözeri, H.; Alveroğlu, E.; Kurtan, U.; Şenel, M.; Baykal, A. Cobalt substituted nickel ferrites via Pechini’s sol–gel citrate route: X-band electromagnetic characterization. J. Supercond. Nov. Magn. 2013, 26, 213–218. [Google Scholar]
  6. Mozaffari, M.; Amighian, J.; Darsheshdar, E. Magnetic and structural studies of nickel-substituted cobalt ferrite nanoparticles, synthesized by the sol–gel method. J. Magn. Magn. Mater. 2014, 350, 19–22. [Google Scholar] [CrossRef]
  7. Yaseneva, P.; Bowker, M.; Hutchings, G. Structural and magnetic properties of Zn-substituted cobalt ferrites prepared by co-precipitation method. Phys. Chem. Chem. Phys. 2011, 13, 18609. [Google Scholar] [CrossRef] [PubMed]
  8. Srinivasamurthy, K.; Angadi, V.; Kubrin, S.; Matteppanavar, S.; Kumar, P.M.; Rudraswamy, B. Evidence of enhanced ferromagnetic nature and hyperfine interaction studies of Ce-Sm doped Co-Ni ferrite nanoparticles for microphone applications. Ceram. Int. 2018, 44, 18878–18885. [Google Scholar] [CrossRef]
  9. Stergiou, C. Magnetic, dielectric and microwave absorption properties of rare earth doped Ni–Co and Ni–Co–Zn spinel ferrites. J. Magn. Magn. Mater. 2017, 426, 629–635. [Google Scholar] [CrossRef]
  10. Boda, N.; Boda, G.; Naidu, K.C.B.; Srinivas, M.; Batoo, K.M.; Ravinder, D.; Reddy, A.P. Effect of rare earth elements on low temperature magnetic properties of Ni and Co-ferrite nanoparticles. J. Magn. Magn. Mater. 2019, 473, 228–235. [Google Scholar] [CrossRef]
  11. Singh, R.K.; Shah, J.; Kotnala, R. Magnetic and dielectric properties of rare earth substituted Ni0.5Zn0.5Fe1.95R0.05O4 (R = Pr, Sm and La) ferrite nanoparticles. Mater. Sci. Eng. B 2016, 210, 64–69. [Google Scholar] [CrossRef]
  12. Akhtar, M.N.; Khan, M.A. Effect of rare earth doping on the structural and magnetic features of nanocrystalline spinel ferrites prepared via sol gel route. J. Magn. Magn. Mater. 2018, 460, 268–277. [Google Scholar] [CrossRef]
  13. Sahariya, J.; Mund, H.S.; Sharma, A.; Dashora, A.; Itou, M.; Sakurai, Y.; Ahuja, B.L. Magnetic properties of NiFe2−xRExO4 (RE=Dy, Gd) using magnetic Compton scattering. J. Magn. Magn. Mater. 2014, 360, 113–117. [Google Scholar] [CrossRef]
  14. Bharathi, K.K.; Noor-A-Alam, M.; Vemuri, R.; Ramana, C.V. Correlation between microstructure, electrical and optical properties of nanocrystalline NiFe1.925 Dy0.075 O4 thin films. RSC Adv. 2012, 2, 941–948. [Google Scholar] [CrossRef]
  15. Almessiere, M.; Slimani, Y.; Güner, S.; Baykal, A.; Ercan, I. Effect of dysprosium substitution on magnetic and structural properties of NiFe2O4 nanoparticles. J. Rare Earths 2019. [Google Scholar] [CrossRef]
  16. Chawla, S.K.; Meena, S.S.; Kaur, P.; Mudsainiyan, R.K.; Yusuf, S.M. Effect of site preferences on structural and magnetic switching Properties of Co–Zr doped strontium hexaferrite SrCoxZrxFe(12-2x)O19. J. Magn. Magn. Mater. 2015, 378, 84–91. [Google Scholar] [CrossRef]
  17. Joseyphus, R.J.; Narayanasamy, A.; Shinoda, K.; Jeyadevan, B.; Tohji, K. Synthesis and magnetic properties of the size-controlled Mn–Zn ferrite nanoparticles by oxidation method. J. Phys. Chem. Solids 2006, 67, 1510–1517. [Google Scholar] [CrossRef]
  18. Chae, K.P.; Kwon, W.H.; Lee, J.-G. Influence of aluminum doping in Li–Co–Ti ferrite. J. Magn. Magn. Mater. 2012, 324, 2701–2705. [Google Scholar] [CrossRef]
  19. Kumar, H.; Sing, J.P.; Srivastava, R.C.; Patel, K.R.; Chae, K.H. Synthesis and characterization of DyxCoFe2−xO4 nanoparticles. Superlattices Microstruct. 2017, 109, 296–306. [Google Scholar] [CrossRef]
  20. Kumar, H.; Srivastava, R.; Singh, J.P.; Negi, P.; Agrawal, H.; Das, D.; Chae, K.H. Structural and magnetic study of dysprosium substituted cobalt ferrite nanoparticles. J. Magn. Magn. Mater. 2016, 401, 16–21. [Google Scholar] [CrossRef]
  21. Morales, M.-B.; Phan, M.-H.; Pal, S.; Frey, N.-A.; Srikanth, H. Particle blocking and carrier fluid freezing effects on the magnetic properties of Fe3O4-based ferrofluids. J. Appl. Phys. 2009, 105, 07B511. [Google Scholar] [CrossRef]
  22. Humbe, A.V.; Kounsalye, J.S.; Shisode, M.V.; Jadhav, K. Rietveld refinement, morphology and superparamagnetism of nanocrystalline Ni0.70−xCuxZn0.30Fe2O4 spinel ferrite. Ceram. Int. 2018, 44, 5466–5472. [Google Scholar] [CrossRef]
  23. Graham, C.D. Introduction to Magnetic Materials; Wiley: Hoboken, NJ, USA, 2009. [Google Scholar]
  24. Bhowmik, R.; Aneeshkumar, K. Low temperature ferromagnetic properties, magnetic field induced spin order and random spin freezing effect in Ni1.5Fe1.5O4 ferrite; prepared at different pH values and annealing temperatures. J. Magn. Magn. Mater. 2018, 460, 177–187. [Google Scholar] [CrossRef]
  25. Ajroudi, L.; Mliki, N.; Bessais, L.; Madigou, V.; Villain, S.; Leroux, C.; Leroux, C. Magnetic, electric and thermal properties of cobalt ferrite nanoparticles. Mater. Res. Bull. 2014, 59, 49–58. [Google Scholar] [CrossRef] [Green Version]
  26. Peddis, D.; Cannas, C.; Piccaluga, G.; Agostinelli, E.; Fiorani, D. Surface spin freezing effects on enhanced saturation magnetization and magnetic anisotropy in CoFe2O4 nanoparticles. Nanotechnology 2010, 21, 125705. [Google Scholar] [CrossRef]
  27. Chen, X.; Bedanta, S.; Petracic, O.; Kleemann, W.; Sahoo, S.; Cardoso, S.; Freitas, P.P.; Freitas, P. Superparamagnetism versus superspin glass behavior in dilute magnetic nanoparticle systems. Physica B 2005, 72, 214436. [Google Scholar] [CrossRef]
  28. Zeb, F.; Sarwer, W.; Nadeem, K.; Kamran, M.; Mumtaz, M.; Krenn, H.; Letofsky-Papst, I. Surface spin-glass in cobalt ferrite nanoparticles dispersed in silica matrix. J. Magn. Magn. Mater. 2016, 407, 241–246. [Google Scholar] [CrossRef]
  29. Almessiere, M.A.; Slimani, Y.; Güngüneş, H.; Baykal, A.; Trukhanov, S.V.; Trukhanov, A.V. Manganese/Yttrium Codoped Strontium Nanohexaferrites: Evaluation of Magnetic Susceptibility and Mossbauer Spectra. Nanomaterials 2019, 9, 24. [Google Scholar] [CrossRef]
  30. Slimani, Y.; Almessiere, M.; Baykal, A. AC susceptibility study of Cu substituted BaFe12O19 nanohexaferrites. Ceram. Int. 2018, 44, 13097–13105. [Google Scholar] [CrossRef]
  31. Kumar, D.; Banerjee, A. A critical examination of magnetic states of La0.5Ba0.5CoO3: Non-Griffiths phase and interacting ferromagnetic-clusters. arXiv 2012, arXiv:1211.4936. [Google Scholar]
  32. Suzuki, M.; Fullem, S.I.; Suzuki, I.S.; Wang, L.; Zhong, C.-J. Observation of superspin-glass behavior inFe3O4nanoparticles. Phys. B 2009, 79, 024418. [Google Scholar] [CrossRef]
  33. Fiorani, D.; Testa, A.M.; Lucari, F.; D’Orazio, F.; Romero, H. Magnetic properties of maghemite nanoparticle systems: Surface anisotropy and interparticle interaction effects. B: Condens. Matter 2002, 320, 122–126. [Google Scholar] [CrossRef]
  34. Rahimi, M.; Kameli, P.; Ranjbar, M.; Salamati, H. The effect of polyvinyl alcohol (PVA) coating on structural, magnetic properties and spin dynamics of Ni0.3Zn0.7Fe2O4 ferrite nanoparticles. J. Magn. Magn. Mater. 2013, 347, 139–145. [Google Scholar] [CrossRef]
  35. Eshraghi, M.; Kameli, P. Structural and Magnetic Properties of La0.9Sr0.1MnO3 Micro and Nanometer-Sized Manganite Samples. J. Mater. Sci. 2014, 3, 1. [Google Scholar]
Figure 1. X-ray diffraction (XRD) powder patterns of prepared nanoparticles (NPs).
Figure 1. X-ray diffraction (XRD) powder patterns of prepared nanoparticles (NPs).
Nanomaterials 09 00820 g001
Figure 2. Elemental mapping, Scanning electron microscope (SEM) images and dispersive X-ray spectroscopy (EDX) of two selected x = 0.01 and 0.05 NPs.
Figure 2. Elemental mapping, Scanning electron microscope (SEM) images and dispersive X-ray spectroscopy (EDX) of two selected x = 0.01 and 0.05 NPs.
Nanomaterials 09 00820 g002
Figure 3. TEM image of selected x = 0.05 nanoferrite (Inset: corresponding selected angle electron diffraction (SAED) pattern).
Figure 3. TEM image of selected x = 0.05 nanoferrite (Inset: corresponding selected angle electron diffraction (SAED) pattern).
Nanomaterials 09 00820 g003
Figure 4. Mössbauer spectra of studied NPs.
Figure 4. Mössbauer spectra of studied NPs.
Nanomaterials 09 00820 g004
Figure 5. MZFC (T) and MFC (T) curves of the three selected NPs.
Figure 5. MZFC (T) and MFC (T) curves of the three selected NPs.
Nanomaterials 09 00820 g005
Figure 6. χ’(T) curves of the prepared (a) NiFe2O4 and (b) NiFe1.97Dy0.03O4 NPs.
Figure 6. χ’(T) curves of the prepared (a) NiFe2O4 and (b) NiFe1.97Dy0.03O4 NPs.
Nanomaterials 09 00820 g006
Figure 7. χ ( T ) curves for (a) NiFe2O4 and (b) NiFe1.97Dy0.03O4 NPs.
Figure 7. χ ( T ) curves for (a) NiFe2O4 and (b) NiFe1.97Dy0.03O4 NPs.
Nanomaterials 09 00820 g007
Figure 8. Plots of (a) ln ( f ) versus 1 / T B and (b) ln ( f ) versus 1 / T s for two selected NiFe2−xDyxO4 NPs where x = 0.00 and 0.03 fit to the Neel–Arrhenius (N–A law) (solid lines).
Figure 8. Plots of (a) ln ( f ) versus 1 / T B and (b) ln ( f ) versus 1 / T s for two selected NiFe2−xDyxO4 NPs where x = 0.00 and 0.03 fit to the Neel–Arrhenius (N–A law) (solid lines).
Nanomaterials 09 00820 g008
Figure 9. Plots of (a) f versus T B and (b) f versus T s for NiFe2−xDyxO4 (x = 0.00 and 0.03) NPs (solid lines present the Vogel–Fulcher (V–F) fit).
Figure 9. Plots of (a) f versus T B and (b) f versus T s for NiFe2−xDyxO4 (x = 0.00 and 0.03) NPs (solid lines present the Vogel–Fulcher (V–F) fit).
Nanomaterials 09 00820 g009
Figure 10. Plots of f versus T s for the two selected x = 0.00 and 0.03 NPs. The solid lines represent the critical slowing down (CSD) fit.
Figure 10. Plots of f versus T s for the two selected x = 0.00 and 0.03 NPs. The solid lines represent the critical slowing down (CSD) fit.
Nanomaterials 09 00820 g010
Table 1. Structural parameters of studied NiDyxFe2−xO4 NPs.
Table 1. Structural parameters of studied NiDyxFe2−xO4 NPs.
xa (Å)V (Å3)DXRD (nm)
0.008.3378579.645134.7
0.018.3397580.033333.1
0.038.3413580.363034.7
0.058.3424580.346330.8
0.078.3473579.382432.3
0.098.3470579.914233.1
0.108.3473580.999824.3
Table 2. Evaluated Mössbauer parameters of studied ferrite NPs (Bhf: hyperfine magnetic field, I.S.: isomer shift, Q.S.: quadrupole splitting, W: line width, and RA: Relative area).
Table 2. Evaluated Mössbauer parameters of studied ferrite NPs (Bhf: hyperfine magnetic field, I.S.: isomer shift, Q.S.: quadrupole splitting, W: line width, and RA: Relative area).
xAssignment of SitesI.S. (±0.002)
(mm/s)
Q.S. (±0.004)
(mm/s)
Bhf (±0.02)
(T)
W (±0.01)
(mm/s)
RA
(%)
Cation Distribution
0.01Sx-A: Fe3+0.3680.00552.0450.21817.298(Ni0.65Fe0.35)A [Ni0.35 Dy0.01Fe1.64]BO4
Sx-B: Fe3+0.4380.09749.4070.40135.575
Sx-B1: Fe3+0.2020.01148.5660.3746.375
Db: Fe+3+0.2260.7170.3990.752
0.03Sx-A: Fe3+0.364052.0090.20517.358(Ni0.66Fe0.34)A [Ni0.34Dy0.03Fe1.63]BO4
Sx-B: Fe3+0.4550.11949.3370.38334.389
Sx-B1: Fe3+0.2040.0948.550.3848.253
0.05Sx-A: Fe3+0.3650.00851.9410.25716.783(Ni0.67Fe0.33)A [Ni0.33Dy0.05Fe1.62]BO4
Sx-B: Fe3+0.4380.07149.3790.39535.528
Sx-B1: Fe3+0.231−0.01848.5590.40247.689
0.07Sx-A: Fe3+0.366−0.00852.0570.24217.866(Ni0.65Fe0.35)A [Ni0.35Dy0.07Fe1.58]BO4
Sx-B: Fe3+0.4640.07249.7370.46625.675
Sx-B1: Fe3+0.218−0.02348.760.42254.952
Db: Fe3+0.3450.6250.7251.5067
0.09Sx-A: Fe3+0.3650.00452.1190.21919.981(Ni0.62Fe0.38)A [Ni0.38Dy0.09Fe1.53]BO4
Sx-B: Fe3+0.4240.06649.6140.44535.589
Sx-B1: Fe3+0.209−0.00548.7660.39445.43
0.10Sx-A: Fe3+0.37−0.00452.1230.23920.331(Ni0.6Fe0.4)A [Ni0.4Dy0.1Fe1.5]BO4
Sx-B: Fe3+0.4410.05349.5440.39731.53
Sx-B1: Fe3+0.198−0.0248.8250.41346.22
Db: Fe3+0.3790.4810.6961.9186
Table 3. Fitting parameters ( τ o , E 0 / k B and ) of prepared NiFe2−xDyxO4 (x = 0.00 and 0.03) NPs estimated using different laws.
Table 3. Fitting parameters ( τ o , E 0 / k B and ) of prepared NiFe2−xDyxO4 (x = 0.00 and 0.03) NPs estimated using different laws.
ModelsParametersValues
Peak (TB)Peak (TS)
x = 0.00x = 0.03x = 0.00x = 0.03
Neel–Arrhenius τ o (s)1.05 × 10−701.68 × 10−321.69 × 10−161.26 × 10−392
E 0 / k B (K)49,08317,4032150184,793
K eff (erg/cm3)3.09 × 1051.09 × 1051.35 × 1041.16 × 106
Vogel–Fulcher τ o (s)5.45 × 10−101.32 × 10−98.00 × 10−76.06 × 10−46
E 0 / k B (K)367.57621.53100.382188.95
T o (K)292.71216.7658.82183.84
K eff (erg/cm3)2.32 × 1033.92 × 1036.34 × 1021.38 × 104
Critical slowing down τ o (s)********2.88 × 10−92.50 × 10−12
T g (K)********58.39205.55
z υ ********5.773.95

Share and Cite

MDPI and ACS Style

Almessiere, M.A.; Slimani, Y.; Güngüneş, H.; Ali, S.; Manikandan, A.; Ercan, I.; Baykal, A.; Trukhanov, A.V. Magnetic Attributes of NiFe2O4 Nanoparticles: Influence of Dysprosium Ions (Dy3+) Substitution. Nanomaterials 2019, 9, 820. https://doi.org/10.3390/nano9060820

AMA Style

Almessiere MA, Slimani Y, Güngüneş H, Ali S, Manikandan A, Ercan I, Baykal A, Trukhanov AV. Magnetic Attributes of NiFe2O4 Nanoparticles: Influence of Dysprosium Ions (Dy3+) Substitution. Nanomaterials. 2019; 9(6):820. https://doi.org/10.3390/nano9060820

Chicago/Turabian Style

Almessiere, Munirah Abdullah, Y. Slimani, H. Güngüneş, S. Ali, A. Manikandan, I. Ercan, A. Baykal, and A.V. Trukhanov. 2019. "Magnetic Attributes of NiFe2O4 Nanoparticles: Influence of Dysprosium Ions (Dy3+) Substitution" Nanomaterials 9, no. 6: 820. https://doi.org/10.3390/nano9060820

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop